Application of molecular orbital theory to transition-metal complexes

Jan 26, 1979 - Application of Molecular Orbital Theoryto Transition-Metal Complexes. 2. Calculation of Enthalpies of Activation for Dissociative Proce...
0 downloads 0 Views 736KB Size
Inorganic Chemistry, Vol. 18, No. 12, 1979 3413

MO Theory and Transition-Metal Complexes Fe(CO)?-, Td, 22321-35-3; CO(CO)F, T d , 14971-27-8; Ni(C0)4, Td, 13463-39-3; Mn(C0)4, D4h, 71564-28-8; MII(CO)~-,D4h, 71564-29-9; F e ( C 0 ) 4 , D4h, 71564-30-2; C O ( C O ) ~ +D4h, , 71564-31-3; Fe(CO)C, D4h, 71564-32-4; cO(co)4, D4hr 71564-33-5; Fe(C0)42-,D4h, 7156434-6; C O ( C O ) ~ -D4h, , 71564-35-7; N i ( C 0 ) 4 , D4h, 71564-36-8; Mn2(CO),,, 10170-69- 1; C O ~ ( C O )1021 ~ , 0-68-1; Fe2H2(C0)s,7 1500-60-2; F e 3 ( C 0 ) , ? , 17685-52-8; H C O ( C O ) ~ ,64519-62-6; HFe(C0)4-, 18716-80-8; H M n ( C 0 ) 5 , 16972-33-1; c i ~ - H ~ F e ( C 022763-20-8; )~, tranr-H2Fe(CO),, 7 1564-37-9; MnCI(CO),, 14100-30-2; MnBr(CO),, 14516-54-2; MnI(CO),, 14879-42-6; M I I ( C H ~ ) ( C O ) ~13601-24-6; , Mn(GeH,)(CO),, 25069-08-3; M ~ ( S I I H , ) ( C O ) ~7,1500-59-9; Mn(PH3)(CO),, 7 1500-58-8; Mn(SH)(CO)5, 59390-73-7; Mn(CH3C= O)(CO),, 13963-91-2; MnC1(C0)4, 715 18-8 1-5; MnBr(CO)4, 71518-82-6; MnI(C0)4, 71518-83-7; Mn(CH3)(CO),, 71518-84-8; Mn(GeH3)(C0)4, 71518-85-9; Mn(SnH3)(C0)4,71518-86-0; Mn(PH,)(CO),, 715 18-87-1; Mn(SH)(C0)4, 7151 8-88-2; Mn(CH3C= O)(CO),, 71518-89-3; V(CO)6, 14024-00-1; V(CO)6-, 20644-87-5; Cr(C0)6, 13007-92-6; Mn(C0)6+, 2133 1-06-6.

References and Notes (1) D. M. P. Mingos, Adu. Organomet. Chem., 15 (1976). (2) For example see (a) D. R. Armstrong, P. G. Perkins, and J. J. P. Stewart, J . Chem. Soc., Dalton Trans., 1972 (1972); (b) D. R. Armstrong, R. Fortune, and P. G. Perkins, J . Catal., 45, 339 (1976); (c) V. D. Sutula, A. P. Zeif, B. I. Popov, and L. I. Chernyauskii, React. Kine?. Catal. Lett., 4, 359 (1976); (d) J. P. Grima, F. Chaplin, and Gr. Kaufmann, J . Organomet. Chem., 124,315 (1977); (e) K. Ohkubo, H. Kanaeda, and K. Tsuchihashi, Bull. Chem. SOC.Jpn., 46,3095 (1973). (3) For example see (a) D. L. Lichtenberger and T. L. Brown, J . Am. Chem. SOC.,100,366 (1978); (b) J. K. Burdett, Inorg. Chem., 16,3013 (1977); (c) D. L. Thorn and R. Hoffmann, J . Am. Chem. SOC.,100, 2079 (1978); (d) H. Berke and R. Hoffmann, ibid., 100,7224 (1978). (4) A. B. Anderson, J . Chem. Phys., 62, 1187 (1975). (5) R. Hoffman, J. Chem. Phys., 39,1397 (1963); R. Hoffman and W. N. Lipscomb, ibid., 36, 2179, 3489 (1962); 37, 2872 (1962). (6) A. B. Anderson, J . Am. Chem. Soc., 100, 1153 (1978), and references cited therein; A. B. Anderson, Inorg. Chem., 15, 2598 (1976). (7) A referee has noted that parameterization which reproduces groundstate geometries is not necessarily applicable to transition-state species (see G. Klopman, P. Andreozzi, A. J. Hopfinger, 0. Kikuchi, and M. J. S. Dewar, J. Am. Chem. Soc., 100, 6267 (1978)). (8) M. Elian and R. Hoffmann, Inorg. Chem., 14, 1058 (1975). (9) J. K. Burdett, J . Chem. Soc., Faraday Trans. 2, 70, 1599 (1974). (10) Ronald J. McKinney and David A. Pensak, Inorg. Chem., following nauer this

rlr . . ....-issue. ._.__

(11) D. G. Schmidling, J . Mole. Strucr., 24, 1 (1975). (12) A. Whitaker and J. W. Jeffrey, Acta Crystallogr., 23, 977 (1967). (13) F. A. Cotton and G. Wilkinson, “Advanced Inorganic Chemistry”, 3rd ed., Interscience, New York, 1972, p 590. (14) J. Demuynck, A. Strich, and A. Veillard, Nouu. J . Chim., 1,217 (1977). (15) P. Jeffrey Hay, J. Am. Chem. Soc., 100, 2411 (1978).

(16) B. Beagley, D. W. J. Cruickshank, P. M. Pinder, A. G. Robiette, and G. M. Sheldrick, Acta Crystallogr., Sect. B, 25, 737 (1969). (17) B. A. Frenz and J. A. Ibers, Inorg. Chem., 11, 1109 (1972). (18) H. Huber, E. P. Kundig, G. A. Ozin, and A. J. Po& J. Am. Chem. SOC., 97, 308 (1975). (19) P. A. Breeze and J. J. Turner, J . Orgunomet. Chem., 44, C7 (1972). (20) R. Perutz and J. J. Turner, Inorg. Chem., 14, 262 (1975). (21) F. A. Cotton, A. Danti, J. S. Waugh, and R. W. Fessenden, J. Chem. Phys., 29, 1427 (1958); R. Bramley, B. N. Figgis, and R. S. Nyholm, Trans. Faraday SOC.,58, 1893 (1962); 0. A. Gansow, A. R. Burke, and W. D. Vernon, J . Am. Chem. SOC.,94, 2552 (1972). (22) M. A. Graham, Ph.D. Thesis, University of Cambridge, 1971. (23) E. P. Kundig and G. A. Ozin, J. Am. Chem. SOC.,96, 3820 (1974), Table I. (24) J. Ladell, B. Post, and I. Faukuchen, Acta. Crystallogr., 5, 795 (1952). (25) G. A. Ozin and A. Vander Voet, Acc. Chem. Res., 6, 313 (1973). (26) 0. Crichton, M. Poliakoff, A. J. Rest, and J. J. Turner, J . Chem. SOC., Dalton Trans., 1321 (1973). (27) M. Poliakoff and J. J. Turner, J . Chem. Soc., Dalton Trans., 2276 (1974). (28) Cited as unpublished work in ref 29. (29) B. Davies, A. McNeish, M. Poliakoff, and J. J. Turner, J . Am. Chem. SOC.,99, 7573 (1977). (30) R. L. DeKock, Inorg. Chem., 10, 1205 (1971). (31) M. Poliakoff, J . Chem. Soc., Dalton Trans., 210 (1974). (32) (a) L. F. Dahl and R. E. Rundle, Acta Crystallogr., 16, 419 (1963); (b) N. I. Gapotchenko, N. V. Alekseev, K. N. Anisimov, N. E. Kolobova, and I. A. Ronova, Zh. Strukt. Khim., 9, 892 (1968); (c) A. Almenningen, G. G. Jacobsen, and H. M. Seip, Acta Chem. Scand., 23, 685 (1969). (33) (a) G. G. Sumner, H. P. Klug, and L. E. Alexander, Acta Crystallogr., 17,732 (1964); (b) J. A. Ibers, J. Organomet. Chem., 14,423 (1968). (34) H. B. Chin, M. B. Smith, R. D. Wilson, and R. Bau, J . Am. Chem. Soc., 96, 5285 (1974). (35) H. M. Powell and R. V. G. Ewens, J. Chem. Soc., 286 (1939). (36) C. H. Wei and L. F. Dahl, J . Am. Chem. Soc., 91, 1351 (1969). (37) A. Anderson, private communication. (38) E. A. McNeill and F. R. Scholer, J . Am. Chem. SOC.,99,6243 (1977). (39) M. B. Smith and R. Bau, J. Am. Chem. SOC.,99, 2388 (1977). (40) S. J. La Placa, W. C. Hamilton, J. A. Ibers, and A. Davison, Inorg. Chem., 8, 1928 (1969). (41) L.n*-\ Vancea and W. A. G. Graham, J . Organornet. Chem., 134, 219 I 3

( l Y I I).

(42) (43) (44) (45)

P. T. Greene and R. F. Bryan, J . Chem. SOC.A , 1559 (1971). H. M. Seip and R. Seip, Acta Chem. Scand., 24, 3431 (1970). H. P. Weber and R. F. Bryan, Chem. Commun., 443 (1966). N. I. Gapotchenko, N. V. Alekseev, A. B. Antonova, K. N. Anisimov, N. E. Kolobova, I. A. Ronova, and T. Yu. Struchkov, J . Organomet. Chem., 23, 525 (1970). (46) E. Clementi and D. L. Raimondi, J . Chem. Phys., 38, 2686 (1963); E. Clementi, D. L. Raimondi, and W. P. Reinhardt, ibid., 46, 1300 (1967). (47) J. W. Richardson, W. C. Nieuwpoort, R. R. Powell, and W. F. Edgell, J . Chem. Phys., 36,1057 (1962); J. W. Richardson, R. R. Powell, and W. C. Nieuwpoort, ibid., 38, 796 (1963). (48) W. Lotz, J . Opt. SOC.Am., 60, 206 (1970).

Contribution No. 2595 from the Central Research and Development Department, Experimental Station, E. I. du Pont de Nemours and Company, Wilmington, Delaware 19898

Application of Molecular Orbital Theory to Transition-Metal Complexes. 2. Calculation of Enthalpies of Activation for Dissociative Processes’ R O N A L D J. M c K I N N E Y and DAVID A. PENSAK*

Received December 21, 1978 Molecular energies calculated with a modified extended Huckel theory are used to calculate enthalpies of activation and/or reaction for a number of dissociative processes involving transition metal carbonyl compounds. Metal-carbonyl, metal-hydride, and metal-metal bond dissociations are examined as well as the interaction of metal hydride with water to give M- H 3 0 + . The effects of a series of ligands, X, on cis CO labilization in MnX(CO)S are examined and compared with similar calculations by other workers by use of the Fenske-Hall method.

+

Introduction In the preceding paper,’ we demonstrated that extended Huckel theory modified by the inclusion of two-body repulsion can successfully reproduce ground-state geometries, including bond lengths, for a number of transition metal carbonyl compounds. In principle, the molecular energies calculated therefrom may be compared to give internal energy ( A E ) of re0020-1669/79/1318-3413$01.00/0 I

action (or activation).2 For example, for eq 1, A E can be calculated by adding the molecular energies of B and C (EB and E c ) and subtracting that of A (EA). The change in A-+B+C (1) AE = E B Ec - E A (2)

+

enthalpy (AH) is more commonly determined experimentally, 0 1979 American Chemical Society

3414 Inorganic Chemistry, Vol. 18, No. 12, 1979

the relationship being as defined in eq 3. For solution chemAH = AE A(nRT) (3) istry under normal experimental conditions the term A(nR7‘) is usually relatively small ( Br known experimental spread of about 15 kcal/mol between the > I, the u and T orbital energies vary in the order C1 C Br former and Mn(SnPh3)(CO)5.11-14The order of effectiveness < I. The latter suggests that iodine should be a better 7 donor of X toward cis labilization as reported by Atwood and than bromine, etc. A comparison of the a overlap between Brown15 is, for the most part, reproduced by our calculations the manganese d and the halogen p orbitals of the complexes and decreases in the order CH3C(0) > S H > CL > Br > I [MnX(CO)5] reveals that C1 > Br > I, strongly suggesting > C H 3 > (PH3)+ > SnH3 > GeH3 > (CO)' > H. The that the degree of overlap is more important than the orbital ordering is consistent with the reported experimental enthalpies energies. This is consistent with photoelectron studies which of activation except that the positions of the germy1 and stannyl indicate that the interaction of the halide orbitals with the ligands are reversed. The positions of the (PH3)+and (CO)+ metal in the ground state decreases in the order C1 > Br > ligands are consistent with qualitative observations of their reactivity and are similarly placed in Atwood and Brown's I.19 Examination of the [Mn(PH3)(CO),]+ case reveals that whereas the p-type orbitals of phosphine are energetically series. l 5 similar to the chlorine p orbitals, their overlap with manganese The complex MnCH3(CO)5 is known not to substitute by d orbitals is considerably reduced because of involvement in carbonyl dissociation but rather by methyl migration to form phosphorus-hydrogen bonding. This is consistent both with an acetyl complex which then decarbonylates, eq 10. Meathe classification of phosphines as poor a donors and with their CH3Mn(CO), + L CH,COMn(CO),L relatively poor labilizing ability. CH3Mn(C0)4L + CO (10) Comparing the calculated activation energies with experisurements for both steps in the case where L = CO have been mental data shows that the former are consistently about 10 obtained by Calderazzo and Cotton" and are A P = 14 and kcal/mol too high. Though some may consider this a large 27 kcal/mol, respectively. If the calculated relationship beerror, it is remarkably small for such an approximate MO tween iodine and methyl is to be trusted, carbonyl dissociation theory. The real value of the technique is in the calculation from MnCH3(CO)5 would require A P = 33-34 kcal/mol, of relative energies of similar systems as is shown by the data thereby explaining the absence of a competitive dissociation presented here. at lower temperatures. Recent kinetic studies on the internal To this point, the calculations described have not distinaromatic metalation of C~S-M~CH,(CO)~P(CH~C~H~)~~~ have guished whether the carbonyl from MnX(CO)5 is cis or trans shown that in this case carbonyl dissociation is the rate-deto X. We have shown only energy differences between optermining step (A# = 32.5 kcal/mol). Calculations on cistimized ground and intermediate states. One approach to MnCH3(C0)4(PH3)and its decarbonylation product give a determining which carbonyl is labilized is to simply remove A P lower than for the pentacarbonyl case (see value in either a cis or the trans carbonyl from the coordination sphere parentheses-Table 11) and very similar to MnI(CO)5. This without letting the resulting intermediate relax to its optimum labilization of MnX(C0)4L compared to MnX(CO)5 has been geometry. Calculating A P by using the "unrelaxed" fragdemonstrated experimentally by Atwood and Brown for X = ments A and B, we find that the ejection of the cis carbonyl Br and L = phosphines, amines, and phosphates.16 The reason is favored over the trans carbonyl by 15 kcal/mol (46.8 kcalfor the lack of a competitive methyl migration reaction path/mol vs. 61.7 kcal/mol). The difference between this cis C O way for the ~ i s - M n c H ~ ( C 0 is ) ~not L clear. ejection value and the one given in Table I1 (41.2 kcal/mol)

-

-

McKinney and Pensak

3416 Inorganic Chemistry, Vol. 48, No. 12, 1979 Table IV

?! OC-Mn

oc'

i,P I

X

A

no.

reaction

calcd A H , kcal/mol

1 2 3 4 5 6 7 8 9 10

CoH(CO), H. + .Co(CO), MnH(C0) -+ H. + ,Mn(CO), FeH,(CO), +H. + .FeH(CO), FeH(CO), + H. + Fe(CO), Mn,(CO),, + 2.Mn(CO), Co,(CO), +2,Co(CO), Fe?H,(CO), -+ 2*FeH(CO), Fc,(CO),, -+3Fe(CO), 2MnH(CO), + H , + Mn,(CO),, 2CoH(CO), -+HZ + Co,(CO), 2FeH,(CO), +H, + Fe,H,(CO), FeH,(CO), + H, + Fe(CO), 3FeH,(CO), + 3H, + Fe,(CO),, CoH(CO), + H,O+H,O' + Co(CO),' FeH,(CO), + H,O + H , O * + FeH(CO),MnH(CO), + H20+H,O' + Mn(CO),FeH(C0); + H,O-+H,O'+ Fc(CO),*-

63.7 64.7 59.8 61.0a 39.6 38.0 10.5 38.2a -14.5 -14.7 t5.0 +16.6a 1-11.5 -3.5 1.2 [5.3] 3.0 [9.3] -5.5 [17.3]

CO OC-Mn-CO

oc'

'I

I

X

B

is due to the energy gained by relaxation of the intermediate to its optimum geometry. The optimized calculated manganese-carbon bond lengths of the ground state do reflect this difference; Le., the trans Mn-C distances optimize 0.03-0.05 A shorter than the cis Mn-C distances, suggesting that the latter bond is somewhat weaker. X-ray crystallographic data for MnCl(CO), show an even larger variation (0.07 A).20 B. M-H Dissociation. The energies calculated for homolytic cleavage of the M-H bond are between 60 and 70 kcal/mol, consistent with the best current estimates for the metal-hydrogen bond strength2' (Table IV, reactions 1-4). C. M-M Dissociation. The energies calculated for metalmetal bond cleavage for four cases are given in Table IV (reactions 5-8). The results for Mn are in excellent agreement with the literature value of about 37 kcal/moLZ2 Though the difference is small, it is reassuring to find the calculated energy of the Co-Co bond is less than that for the Mn-Mn bond, a value consistent with the relative ease of substitution of Co2(CO), compared to Mn2(CO)lo. These reactions are known to take place by initial cleavage of the metal-metal bond.23 Less is known about the reactivity of Fe2H2(CO),; however, the relatively low calculated AH (10.5 kcal/mol) suggests a facile monomer-dimer equilibrium may be important. This is similar to [Fe(q3-C,H,)(C0)3]2which has been experimentally determined at AHdimenzatlon = -1 3 kcal/mol.2A The calculation with Fe3(C0)12should be viewed with caution because we have used the low-spin Fe(C0)4 molecular energy value.7 We also have not explored the dissociation of Fe3(C0)12to give Fe2(C0)*and Fe(CO)+ D. Enthalpies of Hydrogen Elimination from Metal Hydrides. The results from sections B and C (Table IV, reactions 1-8) and a knowledge of the strength of the H2 bond (104.2 kcal/m01)~,can be used to give the thermodynamic relationships found in the middle of Table IV (reactions 9-13). Qualitatively, the results for MnH(CO), and CoH(CO), are correct; when heated or photolyzed under H2, either dimer will form some hydride complex, but under ambient conditions the hydrides will spontaneously decompose to H2and the dimers. FeH,(CO), has been reported to decompose to H2 and the dimer Fe2H2(CO), under reduced pressure.26 On consideration of the positive AH calculated for this reaction, it would be interesting to check the stability of FeH2(C0)4under an atmosphere of hydrogen. It is also noteworthy that what might be expected to be the more likely decomposition pathway for FeH,(C0)4 and conceptually simpler, Le., reductive elimination of H2 followed by clusterification of the resulting [Fe(CO),] fragments to Fe3(C0)12,is calculated to be thermodynamically less favorable and is, in fact, not observed in the laboratory. E. Reaction of M-H + H20. MEHT may also be applied to intermolecular heterolytic cleavage of bonds. One of the simplest cases is the removal of a proton by water from a metal hydride. The energies of H,O and H30+27were therefore calculated, allowing the computation of the results shown in the latter part of Table IV (reactions 14-17). The values shown in parentheses are estimated from known pK, values.26 Though the calculated AH values and the differences between them are incorrect, the relative ordering of the three neutral species is correct. That only the negatively charged [FeH(CO),]- is out of order (and has the wrong sign) is very encouraging and suggests that qualitative studies of ionic solution equilibria may be possible.

11

12 13 14 15 16 17

-+

Using low-spin molecular energy value, see ref 1 for discussion. Estimated from p K , values in ref 26. a

Conclusions The AH values derived from the calculated molecular energies reproduce reasonably accurate energy requirements for metal-carbonyl, metal-hydride, and metal-metal bond dissociation and trends therein for several representative cases. Comparison of our MnX(CO), calculations with the report of Lichtenberger and Brown shows that MEHT gives results qualitatively similar to the more sophisticated nonempirical Fenske-Hall method but more accurately reproduces the relative labilizing capabilities of various ligands. This ability to reproduce relative effects of various substituents on a complex for a specific reaction is being further investigated for a number of systems. In addition, preliminary results suggest that comparison of calculated energies for alternate reaction pathways, substitution in complexes as but one example, may allow the prediction of the dominant pathway and suggest new chemistry to be explored. Acknowledgment. We thank Dr. A. Anderson for providing details of the modified extended Hiickel theory (MEHT) and Dr. F. Van Catledge for helpful discussions. Registry No. MnH(C0)5, 16972-33-1; Mn(C0)6+,21331-06-6; Mn(GeH,)(CO),, 25069-08-3; Mn(SnH,)(C0)5, 7 1500-59-9; Mn(PH3)(C0),+,71500-58-8; Mn(CH,)(CO),, 13601-24-6; MnI(CO),, 14879-42-6; MnBr(CO),, 14516-54-2; MnC1(CO)5, 14100-30-2; Mn(SH)(CO),, 59390-73-7; Mn(CH,C(O))(CO),, 13963-91-2; Mn(CO),', 71 563-56-9; COH(CO)~, 16842-03-8; FeH2(C0)4, 12002-28-7; FeH(CO),, 7 15 16-14-8; MII~(CO)~~, 10170-69-1; C O ~ ( C O 10210)~, 68-1; Fe2H2(CO),, 71500-60-2; F c ~ ( C O ) , ~17685-52-8; , FeH(C0)4-, 187 16-80-8.

References and Notes (1) For part 1, see D. A. Pensak and R. J. McKinney, Inorg. Chem., preceding paper in this issue.

(2) A referee has noted that parameterization which reproduces groundstate geometries is not necessarily applicable to transition-state species (see G. Klopman, P. Andreozzi, A. J. Hopfinger, 0. Kikuchi, and M. J. S. Dewar, J . Am. Chem. Soc., 100, 6267 (1978). (3) D. L. Lichtenberger and T. L. Brown, J . Am. Chem. Soc., 100, 366 (1978). (4) M. B. Hall and R. F. Fenske, Inorg. Chem., 14, 365 (1975). (5) L. Brewer and G. M. Rosenblatt, A h . High Temp. Chem., 2, 1 (1969). (6) L. R. Kangas, R. F. Heck, P. M. Henry, S. Breitschaft, E. M. Thorsteinson, and F. Basolo, J . Am. Chem. Soc., 88, 2334 (1966). (7) See text of preceding paper (ref 1) for discussion of low-spin vs. highspin Fe(C0)4. (8) E. E. Siefert and R. J. Angelici, J . Organomet. Chem., 8, 374 (1967). (9) G. Pajaro, F. Caldera, and R. Ercoli, Gazz. Chim. Ita/., 90, 1486 (1960). (10) We have been informed by D. L. Lichtenberger that very accurate values of AHt could have been obtained through parameterization. They, however, chose to keep the calculations nonempirical.

Inorganic Chemistry, Vol. 18, No. 12, 1979 3417

Properties of Mixed Complexes in Aqueous Solution (11) (12) (13) (14) (15) (16)

(17) (18)

F. Calderazzo and F. A. Cotton, Inorg. Chem., 1, 30 (1962). R. J. Angelici and F. Basolo, J . Am. Chem. SOC.,84, 2495 (1962). G. Jarvinen and H. D. Kaesz, private communication. G. R. Dobson and E. P. Ross, Inorg. Chim. Acta, 5 , 199 (1971). J. D. Atwood and T. L. Brown, J. Am. Chem. Soc., 98, 3160 (1976). J. D. Atwood and T. L. Brown, J . Am. Chem. SOC.,98,3155 (1976). Blaine H. Byers and Theodore L. Brown, J . Organomet. Chem., 127, 181 (1977); J . Am. Chem. Sot., 99, 2527 (1977). These calculations were carried out by using a hydrogen 1 s orbital energy value of -10.0 eV for reasons discussed in ref 1. While use of the more common parameter (-13.6 eV) does change the position of the hydride ligand in the series (about equal with PH,), it must still be considered nonlabilizins. .....~ . ~ - . . . .. -~ ...~ ~ M. B. Hall, J . Am. C h k . Sot., 97, 2057 (1975). P. T. Greene and R. F. Bryan, J . Chem. Soc. A , 1559 (1971). J. A. Comer, Top. Curr. Chem., 71, 71 (1977). L. I. B. Haines, D. Hopgood, and A. J. Po& J . Chem. SOC.A, 421

(23)

(24) (25)

~~~~~

(19) (20) (21) (22)

(26) (27)

(1968); J. L. Hughey, IV, C. P. Anderson, and T. J. Meyer, J . Orgunomet. Chem., 125, C49 (1977). Ronald A. Jackson and Anthony P&, Inorg. Chem., 17,997 (1978), and references cited therein. E. L. Muetterties, B.A. Sosinsky, and K. I. Zamaraev, J . Am. Chem. Soc., 97, 5299 (1975). (a) G. Herzberg, J . Mol. Spectrosc., 33, 147 (1970). (b) Optimization of H2 gave a very reasonable bond length (0.750 .&) when compared with the experimentalvalue (0.746 A). However, the calculated dissociation energy of 153.5 kcal/mol is clearly out of line, and the calculated molecular energy gives unreasonable results when used in the change in enthalpy calculations. The experimental value was found to give very reasonable results and was therefore used. F. A. Cotton and G. Wilkinson, “Advanced Inorganic Chemistry”, 3rd ed., Interscience, New York, 1972, pp 705-6; AH = RT In K. The orbital energy value (-10.00 eV) used for the metal hydrides was used here also.

Contribution from the Istituto Dipartmentale di Chimica e Chimica Industriale, Universitil di Catania, 95 125 Catania, Italy, and the Istituto di Fisica, Universitl di Catania, 95125 Catania, Italy

Thermodynamic and Spectroscopic Properties of Mixed Complexes in Aqueous Solution. Copper(11) Complexes of 2,2’-Bipyridyl and Iminodiacetic or Pyridine-2,6-dicarboxylic Acid FRANCESCO RIGGI,Ib E N R I C O RIZZARELLI,*la R A F F A E L E P. BONOMO,Ia R O S A R I O S I L V I O SAMMARTANO,Ia and GIUSEPPE SIRACUSAIa

Received January 26, 1979 Potentiometric, calorimetric, and EPR investigations on the simple and mixed complexes of copper(I1) with iminodiacetate (ida) or pyridine-2,6-dicarboxylate(dipic) anions and 2,2’-bipyridyl (bpy) were carried out. On the basis of the data obtained, the two ternary complexes Cu(bpy)(dipic) and Cu(bpy)(ida) turned out to have a different coordination number and stereochemistry; in particular the former appears to be octahedrally six-coordinated while the latter is five-coordinated having a square-pyramidal geometry.

Introduction P r e v i o u ~ l ywe , ~ ~reported ~ the thermodynamic and spectroscopic properties of some ternary complexes of Cu” with 2,2’-bipyridyl and some alkane or aromatic dicarboxylic acids. Considering the thermodynamic and spectroscopic (EPR and electronic) data, it was possible to observe that all the mixed complexes had the same structure because of the conformation requirements of 2,2’-bipyridyl ligand, the only exception being the complex with succinate anion, for which a square-pyramidal structure has been proposed. In light of these results we have studied the thermodynamic and spectroscopic (EPR and electronic) parameters of mixed complexes of copper(I1) with 2,2’-bipyridyl (bpy) having the same formula, Cu(bpy)(L), as the above mentioned ones, where L is now a tridentate ligand, namely, iminodiacetate (ida) or pyridine-2,6-dicarboxylate(dipic) anion. It was therefore possible to estimate the effects of the different denticity of the ligands (bpy as bidentate and ida or dipic as tridentate) around the copper(I1) ion and of the presence of different substratum of the tridentate ligands on the thermodynamic properties associated with the formation of ternary complexes, as well as the effects on their structural parameters. The potentiometric (pH and pCu type) and calorimetric measurements were carried out in aqueous solution at 2 5 O C and 0.1 mol dm-3 NaClO,. The thermodynamic and spectroscopic data were compared with those of the corresponding parent complexes measured under the same experimental conditions. We have also obtained all the thermodynamic parameters associated with the formation of the protonated and hydroxo species of the simple complexes of Cu” with the tridentate dicarboxylic acids as well as the AH0 and AS0 values for the formation of the protonated complexes Cu0020-1669179113 18-3417$01 .OO/O

(bpy)(H)(ida) and Cu(bpy)(H)(dipic) by using previously reported potentiometric datae4 Experimental Section Chemicals. The 2,2’-bipyridyl (Erba RPE) was recrystallized from a water-ethanol mixture. The pK value found was in agreement with the value previously determined.5 Iminodiacetic (BDH)and pyridine-2,6-dicarboxylic (Fluka) acids were recrystallized from water. The purity of these acids was checked by titrations with standard C02-freeN a O H and in all cases a value higher than 99% was found. The disodium salts of these acids were prepared according to literature. The preparation of the solid simple and mixed complexes has been described e l ~ e w h e r e . ~All , ~ the solutions were prepared with twicedistilled water and their ionic strength was kept at 0.1 mol dm-3 by addition of NaC104. The standardization of the solutions was carried out as previously described.2 Emf Measurements. The potentiometric measurements were carried out by means of two potentiometers (Amel 232 or Radiometer P H M 52) using a glass electrode (Ingold 201 N S ) or a copper-selective electrode (Amel 201 sens-ion) and a double junction calomel electrode (Orion 90-02-200). In Table I the experimental details of potentiometric titrations are reported. Other details are as previously des~ribed.~~~,*,~ Calorimetric Measurements. The calorimetric measurements were carried out at 25 A 0.001 O C employing a LKB precision calorimeter (Model 8700) and a 100-cm3titration vessel (Model 8726-1). The reproducibility of the system and other details have already been reported.IO Experimental details of the calorimetric titrations are listed in Table 11. Spectroscopic Measurements. First-derivative EPR X-band spectra were recorded with a Varian E-109 instrument equipped with a standard temperature control unit. All measurements reported here were made at 133 K by using quartz sample tubes. The solutions were prepared by dissolving the solid complexes in water-methanol mixtures (1:l ratio) and their concentrations ranged from 1 to 3 mmol dm-j. Field calibration was checked by using polycrystalline di-

0 1979 American Chemical Society